Next Article in Journal
Effect of Temperature on the Morphology and Corrosion Resistance of Modified Boron Nitride Nanosheets Incorporated into Steel Phosphate Coating
Next Article in Special Issue
The Cyclic Stability of the Superelasticity in Quenched and Aged Ni44Fe19Ga27Co10 Single Crystals
Previous Article in Journal
Increasing the Efficiency of Synthetic Iron Production by the Use of New Kit Lining
Previous Article in Special Issue
Terahertz Metamaterial Absorber Based on Ni–Mn–Sn Ferromagnetic Shape Memory Alloy Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetostriction of Heusler Ferromagnetic Alloy, Ni2MnGa0.88Cu0.12, around Martensitic Transition Temperature

1
Mechanics and Robotics Course, Faculty of Advanced Science and Technology, Ryukoku University, Otsu 520-2194, Shiga, Japan
2
Center for Advanced High Magnetic Fields, Graduated School of Science, Osaka University, Toyonaka 560-0043, Osaka, Japan
3
Research Institute for Engineering and Technology, Tohoku Gakuin University, Tagajo 985-8537, Miyagi, Japan
4
Institute for Materials Research, Tohoku University, Sendai 980-8577, Miyagi, Japan
5
Graduate School of Science and Engineering, Yamagata University, Yonezawa 992-8510, Yamagata, Japan
*
Author to whom correspondence should be addressed.
Metals 2023, 13(7), 1185; https://doi.org/10.3390/met13071185
Submission received: 2 June 2023 / Revised: 22 June 2023 / Accepted: 25 June 2023 / Published: 26 June 2023
(This article belongs to the Special Issue Metallic Functional Materials: Development and Applications)

Abstract

:
In this study, magnetostriction measurements were performed on the ferromagnetic Heusler alloy, Ni2MnGa0.88Cu0.12, which is characterized by the occurrence of the martensitic phase and ferromagnetic transitions at the same temperature. In the austenite and martensite phases, the alloy crystallizes in the L21 and D022-like crystal structure, respectively. As the crystal structure changes at the martensitic transition temperature (TM), a large magnetostriction due to the martensitic and ferromagnetic transitions induced by magnetic fields is expected to occur. First, magnetization (M-H) measurements are performed, and metamagnetic transitions are observed in the magnetic field of μ0H = 4 T at 344 K. This result shows that the phase transition was induced by the magnetic field under a constant temperature. Forced magnetostriction measurements (ΔL/L) are then performed under a constant temperature and atmospheric pressure (P = 0.1 MPa). Magnetostriction up to 1300 ppm is observed around TM. The magnetization results and magnetostriction measurements showed the occurrence of the magnetic-field-induced strain from the paramagnetic austenite phase to the ferromagnetic martensite phase. As a reference sample, we measure the magnetostriction of the Ni2MnGa-type (Ni50Mn30Ga20) alloy, which causes the martensite phase transition at TM = 315 K. The measurement of magnetostriction at room temperature (298 K) showed a magnetostriction of 3300 ppm. The magnetostriction of Ni2MnGa0.88Cu0.12 is observed to be one-third that of Ni50Mn30Ga20 but larger than that of Terfenol-D (800 ppm), which is renowned as the giant magnetostriction alloy.

1. Introduction

Ferromagnetic Ni2MnGa-type Heusler alloys are known as highly functional materials in applications such as smart actuators and robotics, which utilizes the magnetic field-induced strain (MFIS) [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30], and magnetic refrigeration, which utilizes the magnetocaloric effect [31,32,33,34,35,36].
In our previous study, we performed thermal expansion, permeability, and magnetization measurements on Ni2MnGa0.88Cu0.12 [37], which is commonly characterized by the occurrence of martensitic and ferromagnetic transitions at the same temperature. Its permeability abruptly changes and shows a clear peak around the martensitic transition temperature, TM. The temperature dependence of the magnetization also shows a clear decrease around TM. The value of TM obtained through linear expansion was 337 K. Furthermore, TM and TR (reverse martensitic transition temperature from martensite phase to austenite phase) increase gradually with increasing magnetic fields.
TM of the ferromagnetic Heusler alloy, Ni2MnGa0.88Cu0.12, in the magnetic field is considered to shift according to the magnetic fields, and it is proportional to the difference between the magnetization of austenite and martensite phases. The shift of TM was estimated as dTM/d0H) = 1.3 K/T. Khovailo et al. [38,39,40] discussed the correlation between the shifts of TM of Ni2 + xMn1−xGa (0 ≤ x ≤ 0.19) using theoretical calculations. The experimental values of this shift corresponded well with the theoretical calculation results. In general, in a magnetic field, the Gibbs free energy is lowered by the Zeeman energy −ΔMB, which enhances the motive force of the martensitic transition. Thus, TM of Ni2MnGa0.88Cu0.12 is considered to have shifted in accordance with the magnetic fields because high magnetic fields are favorable to the ferromagnetic martensitic phases.
The dTM/d(μ0H) of Ni2MnGa0.88Cu0.12 is larger than that of other Ni2MnGa-type Heusler alloys [18,38,40,41], indicating that the magnetic fields considerably affect the martensitic transition of Ni2MnGa0.88Cu0.12 compared to that of other alloys.
Mendonça et al. investigated the magnetizations and magnetostrictions of off-stoichiometric Ni2MnGa-type polycrystalline Ni2Mn0.7Cu0.3Ga0.84Al0.16 alloy [19]. At a temperature slightly higher than TM, a magnetostriction, ΔL/L, caused by the magneto-structural transition demonstrating a significant deformation of 26,000 ppm (2.6%) with metamagnetic transition from the low-magnetic field austenite to the high-magnetic field martensite phase was observed. Martensitic and ferromagnetic transitions occur at the same temperature of TM = 293 K; this property is similar to that of Ni2MnGa0.88Cu0.12 [37]. The shift of TM was estimated as dTM/d(μ0H) = 1 K/T, which is of the same order as that of Ni2MnGa0.88Cu0.12. The temperature dependence of the magnetization M-T curve of Ni2Mn0.7Cu0.3Ga0.84Al0.16 indicated the presence of a metamagnetic-like large hysteresis at 298 K. This, in turn, indicates that at a constant temperature, re-entrant transitions occur between the paramagnetic austenite phase and ferromagnetic martensite phase. In this study, the M-T curve of Ni2MnGa0.88Cu0.12 indicated the metamagnetic-like large hysteresis at 343 and 345 K, which are slightly higher than TM. Moreover, the metamagnetic behavior suggests that Ni2MnGa0.88Cu0.12 can generate a large magnetostriction around TM.
In this study, we analyzed the magnetostriction of Ni2MnGa0.88Cu0.12 and compared its characteristics with those of other MFIS alloys.

2. Materials and Methods

2.1. Materials

The polycrystalline Ni2MnGa0.88Cu0.12 sample was grown by repeated arc melting of the constituent elements, namely 4N Ni, 4N Mn, 4N Cu, and 6N Ga, in an Ar atmosphere. All reaction products were packed in evacuated silica tubes at 1123 K for 72 h and then at 873 K for 24 h before being quenched into water [42,43]. The samples were characterized using X-ray powder diffraction measurements based on Cu-Kα radiation. At room temperature, the crystalline structure was a tetragonal D022-like martensite. In addition, the results of the X-ray diffraction indicated that the tetragonal phase with a D022-like crystal structure coexists with the monoclinic phase of the monoclinic 14M structure (space group: P2/m) at room temperature. The lattice parameters of D022 are aD022 = 0.38920 nm and cD022 = 0.65105 nm. The details of the preparation and characterization of the sample can be found in references [42,43]. As a result of having observed a sample by an optical microscope at 298 K, there are some grains as large as 0.1 mm, which size was smaller than that of polycrystalline Ni2Mn0.7Cu0.3Ga0.84Al0.16 alloy (grain size 0.5 mm) [19].
A Ni2MnGa-type single-crystalline alloy was purchased from Adaptamat Co., Ltd. (Helsinki, Finland). The materials were analyzed using EDS (6010LA, JEOL Co., Ltd., Akishima, Japan) at the Faculty of Advanced Science and Technology, Ryukoku University. The materials included 29.79 at.% Mn, 50.03 at.% Ni, and 20.18 at.% Ga. Therefore, the alloy was transcribed into Ni50Mn30Ga20. An alloy with similar lattice constants and crystal structure as this sample has been presented in [5]. At room temperature, the crystal structure was a 7-layered 7M martensite (which corresponds to the 14M structure and P2/m space group) with lattice constants of a = 0.4219 nm, b = 0.5600 nm, and c = 2.0977 nm. At 324 K, the crystal structure was transformed into cubic-L21 austenite (space group Fm3 m) with a lattice constant of a = 0.5837 nm.

2.2. Methods for Experimental Measurements

Magnetization measurements were performed using the SQUID magnetometer (MPMS-XL7, Quantum Design Co., Ltd., San Diego, CA, USA) at the Center for Advanced High Magnetic Fields, Osaka University. The sample size of Ni2MnGa0.88Cu0.12 for the magnetization measurement was 1.4 × 1.4 × 1.4 mm. The magnetization process M-H measurements were performed at temperatures ranging from 360 K to 330 K. Each measurement point took 30 s. Temperature dependence of the magnetization M-T measurements were performed by warming the sample from 310 K to 400 K, followed by cooling from 400 K to 310 K. The temperature sweep ratio was 0.016 K/s. Each measurement point took 60 s.
Magnetostriction measurements were carried out using strain gauges KFR-02-120-C1-16 (Kyowa Dengyo Co., Ltd., Chofu, Japan) under atmospheric pressure and without applying any pre-stress. The base size of this strain gauge is 2.5 mm (parallel to the grid) × 2.2 mm. The electrical resistivity of the strain gauges was measured using the four-probe method and was measured by means of 2000/J digital voltmeter (Keithley Co., Ltd., Cleveland, OH, USA). This digital voltmeter was connected to a CF-FX1 Laptop PC (Panasonic Co., Ltd., Kadoma, Japan) via a GPIB cable, and measured by means of Basic 98 program. The sample size of Ni2MnGa0.88Cu0.12 is 3.5 × 3.0 × 3.0 mm. The magnetostrictions of Ni2MnGa0.88Cu0.12 were measured parallel to the applied magnetic field (ΔL/L)//H. The sweep speed of the magnetic field during the magnetostriction measurement was 0.0125 T/s. Each magnetostriction measurement at a specific temperature was measured after heating the sample to 370 K, which is higher than TR, and then maintaining a constant temperature. In order to estimate the error of the strain gauge, the strain gauge was fixed on Cu plate, and the magnetoresistance of the strain gauge was measured in a magnetic field up to 10 T. The error was estimated to be ±1 ppm at 5 T and ±3 ppm at 10 T.
The forced magnetostriction measurements for Ni2MnGa0.88Cu0.12 were performed using a helium-free superconducting magnet up to 6 T at the Center for Advanced High Magnetic Field Science, Osaka University, and up to 9.8 T at the High Field Laboratory for Superconducting Materials, Institute for Materials Research, Tohoku University.
The forced magnetostriction measurement for Ni50Mn30Ga20 was performed using a four-prove strain gauge method up to 1.5 T with a water-cooled electromagnet (Tamagawa Seisakusho Co., Ltd., Sendai, Japan) at the Faculty of Advanced Science and Technology, Ryukoku University. The sample size was 3.5 × 2.5 × 1.0 mm, and the longitudinal direction of the sample was parallel to the c-axis. The magnetostriction was measured along the c-axis in a magnetic field applied parallel to the c-axis, (ΔL/L)//H//c.

3. Results and Discussions

3.1. Temperature and Magnetic Field Dependency of the Magnetization

Figure 1 shows the temperature dependence of the magnetization for Ni2MnGa0.88Cu0.12 in a magnetic field of μ0H = 0.1 T. The low- and high-temperature phases below and above 340 K indicate the ferromagnetic martensite and paramagnetic austenite phases, respectively. This is consistent with the results of previous studies [37,42,43]. The reason of the dip of the magnetization M-T curve around 325 K in cooling process, and the difference of the M-T curve in cooling and heating processes shown in Figure 1, can be attributed to the alignment of magnetic moments of Mn and Ni atoms parallel to the magnetic field just below TM and the rearrangement of the 14M and/or DO22 tetragonal lattices by the magnetic moments.
The values of dM/dT shown in Figure 2 indicate magnetization with respect to temperature. The peak temperature of dM/dT corresponds to the transition temperature, indicating the martensitic transition temperature of TM = 338 K (austenite phase to martensite phase). The reverse martensitic transition temperature is almost the same as TM, with a value of 337 K, which was obtained through linear expansion in our previous study [37]. Based on the heating process of dM/dT, TR = 343 K (martensite phase to austenite phase). From the dM/dT results shown in Figure 2, the martensitic transition start temperature TMs, the martensitic transition finish temperature TMf, the reverse martensitic transition start temperature TRs, and the reverse martensitic transition finish temperature TRf, was obtained as TMs = 349 K, TMf = 335 K, TRs = 340 K, TRf = 353 K. The martensitic temperature TM as 338 K was obtained from the peak temperature of dM/dT during the cooling process. The reverse martensitic transition temperature TR was obtained from the peak temperature of dM/dT in the heating process.
In Figure 3, the magnetic field dependence of the magnetization for Ni2MnGa0.88Cu0.12 at a constant temperature is shown. It can be observed that the martensite phase has larger magnetization than that of the austenite phase. Metamagnetic-like hysteresis was observed around 340 K. This indicates that, at the constant temperature, transition from low-magnetic field paramagnetic austenite phase to high-magnetic field ferromagnetic martensite phase occurred. This is similar to the re-entrant transition between paramagnetic austenite phase and ferromagnetic martensite phase as shown in Ni2Mn0.7Cu0.3Ga0.84Al0.16 polycrystalline alloy [19].
The magnetostriction was measured to investigate the relation between the structural phase transition and magnetic property.

3.2. Forced Magnetostriction

The forced magnetostriction measurements for Ni2MnGa0.88Cu0.12 were performed from a higher temperature to a lower temperature compared to the martensite transition temperature.
Figure 4 shows the magnetic field dependence of the forced magnetostriction, ΔL/L, up to 6 T. First, we measured it at 352 K, which is higher than TM and then at a constant temperature, while gradually cooling the sample. For the y-axis (ΔL/L), an offset is applied.
The magnetostriction increases when the sample is cooled from 352 K, and reaches its maximum at 1300 ppm near 343 K. Furthermore, the magnetostriction decreases when the sample is cooled. These results indicate that large magnetostriction was observed at a slightly higher temperature than the martensite transition temperature, TM = 338 K. Furthermore, large hysteresis was observed at 345 and 343 K, which could be equivalent to the hysteresis of magnetization; in addition, re-entrant transition occurred between the paramagnetic austenite and ferromagnetic martensite phases.
As for Ni2MnGa alloys, the rearrangement of the martensite variants is responsible for a large part of the magnetic field-induced strain. The large plastic strain observed between 341 K and 345 K indicates that this effect plays a significant and important role. In Figure 4, the magnetostriction curve at 352 K and 350 K, in the fully austenite phase, which has a temperature lower than TMs = 349 K, indicates a small hysteresis. At 348 K, the hysteresis of the magnetostriction curve is larger than that at 350 K. This indicates that the temperature was slightly lower than TMs, and thus the martensitic transition took place. At 345 K, the largest hysteresis was found. At this temperature, we expect that the austenite parent phase at low magnetic fields would change into the martensite phase. The result of thermal strain in the cooling process indicates a contraction of 1300 ppm (0.13%) with a martensitic transition around TM [37]. The shrinkage value of the magnetostriction curve is comparable to the contraction value of the thermal strain result observed in our former study. For the isothermal magnetization process at 344 K, as shown in Figure 3, the M-H curve for decreasing magnetic field returns to the initial phase before the field reaches zero, suggesting a considerable recovery of the austenite phase. Therefore, a large magnetic field-induced strain with a large hysteresis and significant austenite recovery at 345 K was observed. These results are consistent with those of Ni2.15Mn0.70Cr0.15Ga polycrystal as shown in reference [20]. From the magnetization and the magnetostriction results, the magnetic driving force is expected to be stronger than the potential barrier for the martensitic transition from the austenite phase to the martensite phase.
At 341 K, which is higher than TMf = 335 K, there are still some austenite regions in zero magnetic field. Above 4 T, the strain value was almost constant, which indicates that fully martensitic state was realized, and the variants were fully aligned in a high magnetic field. As the magnetic field-induced martensitic transition occurred at high magnetic fields, it is conceived that the recovery strain became small as it is difficult to exceed the potential barrier in low magnetic fields.
At 338 K, a nearly fully martensite state was realized even at zero magnetic field, resulting in small magnetostriction.
Figure 5 shows the magnetic field dependence of the forced magnetostriction, ΔL/L, up to 9.8 T. At 345 K, the re-entrant transition is clearly identified. At 347 K, the transition shrank almost linearly with increasing magnetic fields, demonstrating hysteresis. At 345 K, the sample also shrank almost linearly with increasing magnetic fields. The shrinkage decreased beyond 8 T. When the magnetic field decreased from 9.8 T, the length of the sample remained almost the same, and below 6 T, the sample gradually lengthened. At 1.5 T, a soft kink was observed in ΔL/L. The result of magnetostriction at 345 K indicates that the magnetic-field-induced transition from the paramagnetic austenite to ferromagnetic martensite phase occurred in increasing magnetic field process.
In [42] and [43], regarding Ni2MnGa1-xCux, the magnetic field was conceived to shrink because of the x dependence of Cu concentration of the unit-cell volume within the interval 0 x 0.25 at room temperature. The unit-cell volume decreases with increasing x. Ni2MnGa1−xCux in the concentration range of 0 ≤ x ≤ 0.07 crystallize in the L21 cubic structure at room temperature. The lattice parameter aL21 with x = 0.02 (L21 cubic structure) was 0.58206 nm. For 0.10 ≤ x ≤ 0.25, the sample crystallizes as a D022-like martensite structure, as mentioned in Section 2.1. At x = 0.12, the unit-cell volume defined through X-ray diffraction was smaller than that of the extrapolation line at 0 ≤ x ≤ 0.07 (L21 cubic structure) of the order of 0.200% (2000 ppm). The linear strain of a polycrystal is one-third of the volume strain [44]. Therefore, 670-ppm shrinkage is expected because of the austenite phase transition to the martensite phase. Moreover, in the martensite phase, the magnetic easy axis is aD022, and a large shrinkage was observed with the applied magnetic field parallel to the easy axis [15]. Therefore, negative magnetostriction could be observed around TM for this study.

3.3. Comparation with Other Ferromagnetic Magneto-Structural Alloys

Now, we compare the magneto-structural properties of Ni2MnGa0.88Cu0.12 with other Ni2MnGa-type ferromagnetic alloys, i.e., Ni2MnGa [17], Ni41Co9Mn31.5Ga18.5 [31], Ni2Mn0.7Cu0.3Ga0.84Al0.16 [19], and Ni2.15Mn0.70Cr0.15Ga [20] polycrystalline samples, as well as Ni45Co5Mn36.7In13.3 [7] and Ni50Mn30Ga20 [5] single crystalline samples.
Table 1 provides information on the crystal structure in the martensite phase, volume change from the martensite phase to the austenite phase, and magnetostriction value of varioys Ni2MnGa-type Heusler alloys.
Sakon et al. investigated the magnetostriction of polycrystalline sample of Ni2MnGa [17]. The magnetostriction was measured at T = 185 K in the martensite phase, where the temperature was just below TM = 193 K. The magnetostriction was measured parallel ( Δ L / L )// and perpendicular ( Δ L / L ) to the magnetic field. Furthermore, the obtained magnetostriction values were ( Δ L / L )// = −780 ppm and ( Δ L / L ) = 260 ppm. The volume magnetostriction Δ V / V can be obtained as follows [44,45,46]:
Δ V / V = ( Δ L / L ) / / + 2   ( Δ L / L )
The obtained Δ V / V was −260 ppm. Singh et al. [45] obtained the temperature dependencies of the lattice constant and unit-cell volume based on a high-resolution synchrotron X-ray powder diffraction. The volume change from a L 2 1 cubic structure to a monoclinic 14M structure was −330 ppm. Volume magnetostriction Δ V / V was 80% of the volume change of the L21   14M martensite transition.
Ni41Co9Mn31.5Ga18.5 has a TM of 320 K and a Curie temperature, TC of 468 K [31]. The L21 cubic ferromagnetic austenite structure and tetragonal ferrimagnetic martensite structure were obtained as magnetic and crystalline structure, respectively. Thermo-magnetic M–T curves indicated that at 0.5 T, a sharp magnetization jump corresponding to the thermal martensitic transition temperatures [martensite → austenite TR = 350 K (heating process) and austenite → martensite TM = 320 K (cooling process)] was observed at around room temperature. The TM decreased with increasing magnetic field at the ratio of dTM/d(μ0H) = −12.6 K/T. The absolute value of dTM/d(μ0H) of Ni41Co9Mn31.5Ga18.5 is 10 times larger than that of Ni2MnGa0.88Cu0.12. Below TM, metamagnetic transitions were observed for the magnetization processes (M-T curves). In addition, magnetostrictions larger than 1000 ppm were observed at the temperature of 150 K   T 290 K below TM, corresponding to a metamagnetic transition. At 290 K, the total value of magnetostriction was 1900 ppm. Assuming isotropic magnetostriction for this polycrystalline sample, the total volume change can be estimated as 5700 ppm [31]. These results indicate that below TM, a reverse martensite transition occurred from the ferrimagnetic martensite phase to the ferromagnetic austenite phase with the increasing magnetic field. The experimental and theoretical results show that the total energy of the ferromagnetic state is close to that of the ferrimagnetic martensite state, and the magnetic-field-induced ferromagnetic austenite phase is unstable in the ratio of the tetragonality (c/a) for 1.0     c/a  1.1 . Therefore, metamagnetic transitions and large magnetostrictions were observed in the magnetic field.
A single crystal of Ni45Co5Mn36.7In13.3 indicates a large MFIS of 30,000 ppm (3.0%) at 298 K when a compressive pre-strain of approximately 3% was applied in the direction plotted with a filled circle in the stereographic triangle, as shown in Figure 4 of Ref. [7]. The magnetic field was applied parallel to the compressive axis of the sample. The TM was 290 K and TC was 382 K. Below this TM, the M-T curve indicated a clear metamagnetic transition behavior. The TM decreased with increasing magnetic field at the ratio of dTM/d(m0H) = −7 K/T, and an L21 cubic structure was achieved in austenite phase, and a monoclinic 14 M layered structure was achieved in the martensite phase. Below TM, the MFIS was observed to occur during the metamagnetic transition from the ferrimagnetic martensite phase to ferromagnetic austenite phase. The strain from the L21 cubic structure to monoclinic 14 M layered structure is significant, exceeding a few percent. For example, Sozinov et al. experimentally studied the MFIS of a single-variant sample of an orthorhombic seven-layered phase in the Ni48.8Mn29.7Ga21.5 single crystalline alloy at 300 K, measuring it perpendicular to the magnetic field applied along the [100] direction [3]. They observed a giant MFIS of approximately 95,000 ppm (9.5%) at an ambient temperature in a magnetic field of less than 1 T in the Ni2MnGa orthorhombic seven-layered (7M structure, which corresponds to 14M structure) martensite phase.
On the contrary, the crystal structure of Ni41Co9Mn31.5Ga18.5 is not a monoclinic 14 M structure but a tetragonal structure, and the volume change due to the martensitic transition is 5700 ppm. Therefore, a moderate magnitude magnetostriction was observed.
In the case of Ni2MnGa0.88Cu0.12, the value of the magnetostriction (1300 ppm) is the same order as that of Ni41Co9Mn31.5Ga18.5 (1900 ppm). The tetragonal phase of the D022-like crystal structure coexists with the monoclinic phase of the monoclinic 14M structure (space group: P2/m) in the martensite phase at room temperature. The volume change from the L21 cubic structure to monoclinic 14M structure was −330 ppm for Ni2MnGa [45]. Therefore, −1300 ppm magnetostriction, which is smaller than that of Ni41Co9Mn31.5Ga18.5, the crystal structure of which is a tetragonal martensite, is suitable for the Ni2MnGa0.88Cu0.12 polycrystalline alloy.
As mentioned earlier, for Ni2Mn0.7Cu0.3Ga0.84Al0.16, TM = 293 K and the martensitic and ferromagnetic transitions occur at the same temperature, as confirmed by the magnetization and magnetostriction measurements [19]. In addition, the results of X-ray powder diffraction measurement indicated that the crystal structure was L21 cubic austenite at 312 K. Moreover, at 298 K, which is slightly higher than TM = 293 K, the austenitic L21 cubic structure and martensitic L10 non-modulated tetragonal structure were observed to coexist. Finally, below 257 K, the L10 structure was achieved.
The unit-cell volume of Ni2Mn0.7Cu0.3Ga0.84Al0.16 calculated by the lattice parameters showed the volume difference between the austenite L21 phases and martensite L10 phase as approximately −55,000 ppm (−5.5%). Therefore, a giant magnetostriction was expected around −18,000 ppm (−1.8%) for the polycrystalline sample. However, the experimental results showed magnetostriction of ΔL/L = −26,000 ppm (−2.6%.), which is larger than the expected value. Therefore, the effect of the alignment of the variant of martensite must be considered. The magnetostriction is irreversible during the experimental temperatures of 297.4 K T 304 K. Partial recovery was observed only at 305 K.
In the case of Ni2MnGa0.88Cu0.12, the value of the magnetostriction was one-tenth that of Ni2Mn0.7Cu0.3Ga0.84Al0.16. However, an 85% recovery strain was observed at 345 K. The existence of recovery strain at atmospheric pressure is advantageous for the alloy’s application in sensors and actuators.
Mendonça et al. [20] performed the magnetostriction measurements on the Ni2.15Mn0.70Cr0.15Ga polycrystalline alloy with TM = 299 K. In addition, the martensitic and ferromagnetic transitions occurred at the same temperature, as confirmed by the magnetization and magnetostriction measurements. The resulting crystal structure in the parent austenite phase was the L21 cubic structure. The martensite phase is an L10 non-modulated tetragonal structure. The unit-cell volume of Ni2Mn0.7Cu0.3Ga0.84Al0.16 calculated from the lattice parameters showed that the volume difference between the austenite L21 phases and martensite L10 phase is approximately −30,000 ppm (−3.0%). The magnetostriction results indicated a reversible strain around −8100 ppm (−0.81%) under 0–9 T, which is slightly lesser than one-third the volume difference (−10,000 ppm, −1.0%). Mendonça et al. [20] overcame the weak point of large hysteresis in the magnetostriction of Ni2Mn0.7Cu0.3Ga0.84Al0.16. The characteristic of a magneto-structural transformation at room temperature with low hysteresis is advantageous for the alloy’s application in functional magnetic materials. In this experimental study, the value of the magnetostriction of Ni2MnGa0.88Cu0.12 is one-sixth of that of Ni2.15Mn0.70Cr0.15Ga. However, it is advantageous for its industrial application and possesses smaller magnetostriction than that of Ni2.15Mn0.70Cr0.15Ga.
As a reference sample, we measured the magnetostriction of the Ni2MnGa-type (Ni50Mn30Ga20) alloy manufactured by Adaptamat Co., Ltd. [4,5], which is shown in Figure 6. This alloy causes martensite phase transition at TM = 315 K. Magnetostriction of Ni50Mn30Ga20 was measured parallel to the magnetic field, which was parallel to the c-axis. Magnetostriction of −3300 ppm was observed at an atmospheric pressure and 298 K. Almost 100% recovery strain was observed. The value of the magnetostriction of Ni2MnGa0.88Cu0.12 was one-third that of Ni50Mn30Ga20. However, the value of the magnetostriction of Ni2MnGa0.88Cu0.12 is larger than that of Terfenol-D (800 ppm), which is renowned as the most commonly used magnetostrictive alloy [47,48].
As mentioned above, a shrinkage of −3300 ppm was observed at 298 K in the martensite phase. This result is comparable to magnetostriction of Ni2MnGa single crystal at atmospheric pressure (unstressed crystal) measured by Ullakko et al. [1]. The martensitic transition temperature, TM, was 275 K. In the martensite phase at 265 K, a magnetostriction value of the [001] axis (c-axis in a martensite phase) strain in response to magnetic field applied along c-axis was—1000 ppm in an applied magnetic field of 1 T. Ullakko et al. mentioned that the strain occurs fully within the martensitic phase and is due to the motion of twin boundaries. The magnetostriction is only a small fraction of the lattice constant change (∆c/c = 6.56% = 65,600 ppm). They also mentioned that this small strain is caused by the strain accommodation through different twin variant orientations. In the martensite phase, applying the magnetic field to the single variant material causes the other twin variants to appear and grow [3,4]. When magnetic field strength increases the boundaries between twins move, and the preferentially oriented twin variants grow at the expense of the other twin variants. It is interesting that the magnetostriction of Ni50Mn30Ga20 (this study) and Ni2MnGa [1] was almost fully reversible. Currently, we cannot explain why the recovery of strain occurred in the martensite phase. It is conceived that the rearrangements of the twin variants would occur when increasing a magnetic field, and afterwards, during the decreasing magnetic field process, elastic recovery may act and come back to the original state. However, the exact reason for this recovery is currently not clear. Recently, Wu et al. simulated the ferromagnetic domain structure and martensite variant microstructure of Ni2MnGa-type shape-memory alloy [12]. They found −0.28% (−2800 ppm) magnetic field induced strain and also found 100% recovery strain at 285 K below TM = 300 K. This value is comparable with our result of Ni50Mn30Ga20 alloy. We retrieve this model helps explanation of the recovery strain.

4. Conclusions

In this study, magnetostriction measurements were performed on the ferromagnetic Heusler alloy Ni2MnGa0.88Cu0.12. One characteristic of this alloy is that its martensitic and ferromagnetic transitions are caused at the same temperature. In the austenite and martensite phases, the alloy crystallizes with L21 and D022-like crystal structures, respectively. Metamagnetic transition was observed in the magnetic field of μ0H = 4 T at 344 K. This result showed that the occurrence of phase transition was induced by the magnetic field under constant temperature. Forced magnetostriction measurements (ΔL/L) were performed under constant temperature and at an atmospheric pressure of P = 0.1 MPa. Magnetostriction up to 1300 ppm was observed around the TM. The magnetization and magnetostriction measurement results showed that the magnetic-field-induced strain from the paramagnetic austenite phase to ferromagnetic martensite phase occurred. In the case of Ni2MnGa0.88Cu0.12, the value of the magnetostriction was one-tenth that of Ni2Mn0.7Cu0.3Ga0.84Al0.16, which has a tetragonal L10 martensite structure. However, 85% recovery strain was observed at 345 K. The value of the magnetostriction of Ni2MnGa0.88Cu0.12 was one-sixth that of Ni2.15Mn0.70Cr0.15Ga, which has low hysteresis and reversible strain in magnetic fields. The existence of a recovery strain at atmospheric pressure is advantageous to the alloy’s application to sensors and actuators.
For reference, we measured the magnetostriction of the Ni2MnGa-type (Ni50Mn30Ga20) alloy made in Adaptamat Co., Ltd. The magnetostriction of −3300 ppm was observed at 298 K, and almost 100% recovery strain was observed. The value of the magnetostriction of Ni2MnGa0.88Cu0.12 was one-third that of Ni50Mn30Ga20. However, the value of the magnetostriction of Ni2MnGa0.88Cu0.12 is larger than that of Terfenol-D (800 ppm).

Author Contributions

Y.A. and T.K. prepared the samples; T.S., H.N., Y.N., M.H. and T.K. conceived and designed the experiments; K.M., T.K., H.N., Y.N., M.H. and T.S. performed the experiments; K.M. and T.S. analyzed the data; T.S., K.M., T.K., Y.N., M.H., H.N. and Y.A. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

The authors express their gratitude to K. Endo for their contribution to sample preparation. This research was carried out at the Center for Advanced High Magnetic Field Science in Osaka University under the Visiting Researchers’ Program of the Institute for Solid State Physics, the University of Tokyo and at the High Field Laboratory for Superconducting Materials, Institute for Materials Research, Tohoku University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ullakko, K.; Huang, J.K.; Kantner, C.; O’Handley, R.C.; Kokorin, V.V. Large magnetic-field-induced strains in Ni2MnGa single crystals. Appl. Phys. Lett. 1996, 69, 1966–1968. [Google Scholar] [CrossRef]
  2. Ullakko, K.; Huang, J.K.; Kokorin, V.V.; O’Handley, R.C. Magnetically controlled shape memory effect in Ni2MnGa intermetallics. Scr. Mater. 1997, 36, 1133–1138. [Google Scholar] [CrossRef]
  3. Sozinov, A.; Likhachev, A.A.; Lanska, N.; Ullakko, K. Giant magnetic-field-induced strain in NiMnGa seven-layered martensitic phase. Appl. Phys. Lett. 2002, 80, 1746–1748. [Google Scholar] [CrossRef]
  4. Tellinen, J.; Suorsa, I.; Jääskeläinen, A.; Aaltio, I.; Ullakko, K. Basic properties of magnetic shape memory actuators. In Proceedings of the 8th International Conference ACTUATOR, Bremen, Germany, 10–12 June 2002; pp. 566–569. [Google Scholar]
  5. Ge, Y.; Söderberg, O.; Lanska, N.; Sozinov, A.; Ullakko, K.; Lindroos, V.K. Crystal structure of three Ni-Mn-Ga alloys in powder and bulk materials. J. Phys. IV 2003, 112, 921–924. [Google Scholar] [CrossRef]
  6. Chernenko, V.A.; L’vov, V.A. Magnetoelastic nature of ferromagnetic shape memory effect. Mater. Sci. Forum 2008, 583, 1–20. [Google Scholar] [CrossRef] [Green Version]
  7. Kainuma, R.; Imano, Y.; Ito, W.; Sutou, Y.; Morito, H.; Okamoto, S.; Kitakami, O.; Oikawa, K.; Fujita, A.; Kanomata, T.; et al. Magnetic-field-induced shape recovery by reverse phase transformation. Nature 2006, 439, 957–960. [Google Scholar] [CrossRef]
  8. Sakon, T.; Sasaki, K.; Numakura, D.; Abe, M.; Nojiri, H.; Adachi, Y.; Kanomata, T. Magnetic field-induced transition in co-doped Ni41Co9Mn31.5Ga18.5 heusler alloy. Mater. Trans. 2013, 54, 9–13. [Google Scholar] [CrossRef] [Green Version]
  9. Karaca, H.E.; Karaman, I.; Basaran, B.; Ren, Y.; Chumlyakov, Y.I.; Maier, H.J. Magnetic field-induced phase transformation in NiMnCoIn magnetic shape-memory alloys—A new actuation mechanism with large work output. Adv. Funct. Mater. 2009, 19, 983–998. [Google Scholar] [CrossRef]
  10. Kainuma, R.; Oikawa, K.; Ito, W.; Sutou, Y.; Kanomata, T.; Ishida, K. Metamagnetic shape memory effect in NiMn-based Heusler-type alloys. J. Mater. Chem. 2008, 18, 1837. [Google Scholar] [CrossRef]
  11. O’Handley, R.C.; Murray, S.J.; Marioni, M.; Nembach, H.; Allen, S.M. Phenomenology of giant magnetic-field-induced strain in ferromagnetic shape memory materials. J. Appl. Phys. 2000, 87, 4712–4717. [Google Scholar] [CrossRef]
  12. Wu, P.; Liang, Y. Enhanced reversible magnetic-field-induced strain in Ni-Mn-Ga Alloy. Metals 2021, 11, 2017. [Google Scholar] [CrossRef]
  13. Kumar, A.S.; Seshubai, V. 0.7% magnetic field induced strain in polycrystalline Ni50Mn29Ga21 ferromagnetic shape memory alloy. Int. J. Innov. Res. Sci. Eng. Technol. 2013, 2, 4226–4232. [Google Scholar]
  14. Mennerich, C.; Wendler, F.; Jainta, M.; Nestler, B. Rearrangement of martensitic variants in Ni2MnGa studied with the phase-field method. Eur. Phys. J. B 2013, 86, 171. [Google Scholar] [CrossRef]
  15. Okamoto, N.; Fukuda, T.; Kakeshita, T.; Takeuchi, T.; Kishino, K. Rearrangement of variants in Ni2MnGa under magnetic field. Sci. Technol. Adv. Mater. 2004, 5, 29–34. [Google Scholar] [CrossRef]
  16. Sakon, T.; Fujimoto, N.; Kanomata, T.; Adachi, Y. Magnetostriction of Ni2Mn1−xCrxGa heusler alloys. Metals 2017, 7, 410. [Google Scholar] [CrossRef] [Green Version]
  17. Sakon, T.; Yamasaki, Y.; Kodama, H.; Kanomata, T.; Nojiri, H.; Adachi, Y. The characteristic properties of magnetostriction and magneto-volume effects of Ni2MnGa-type ferromagnetic heusler alloys. Materials 2019, 12, 3655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Sakon, T.; Adach, Y.; Kanomata, T. Magneto-structural properties of Ni2MnGa ferromagnetic shape memory alloy in magnetic fields. Metals 2013, 3, 202–224. [Google Scholar] [CrossRef]
  19. Mendonça, A.A.; Jurado, J.F.; Stuard, S.J.; Silva, L.E.L.; Eslava, G.G.; Cohen, L.F.; Ghivelder, L.; Gomes, A.M. Giant magnetic-field-induced strain in Ni2MnGa-based polycrystal. J. Alloys Compd. 2018, 738, 509–514. [Google Scholar] [CrossRef]
  20. Mendonça, A.A.; Ghivelder, L.; Bernardo, P.L.; Cohen, L.F.; Gomes, A.M. Low hysteretic magnetostructural transformation in Cr-doped Ni-Mn-Ga Heusler alloy. J. Alloys Compd. 2023, 938, 168444. [Google Scholar] [CrossRef]
  21. Sofronie, M.; Tolea, F.; Tolea, M.; Popescu, B.; Valeanu, M. Magnetic and magnetostrictive properties of the ternary Fe67.5Pd30.5Ga2 ferromagnetic shape memory ribbons. J. Phys. Chem. Sol. 2020, 142, 109446. [Google Scholar] [CrossRef]
  22. Mahfouzi, M.; Carman, G.P.; Kioussis, N. Magnetoelastic and magnetostrictive properties of Co2XAl Heusler compounds. Phys. Rev. B 2020, 102, 094401. [Google Scholar] [CrossRef]
  23. Sofronie, M.; Tolea, M.; Popescu, B.; Enculescu, M.; Tolea, F. Magnetic and Magnetostrictive Properties of Ni50Mn20Ga27Cu3 Rapidly Quenched Ribbons. Materials 2021, 14, 5126. [Google Scholar] [CrossRef]
  24. Liu, K.; Ma, S.; Zhang, Y.; Zeng, H.; Yu, G.; Luo, X.; Chena, C.; Rehman, S.U.; Hu, Y.; Zhong, Z. Magnetic-field-driven reverse martensitic transformation with multiple magneto-responsive effects by manipulating magnetic ordering in Fe-doped Co-V-Ga Heusler alloys. J. Mater. Sci. Technol. 2020, 58, 145–154. [Google Scholar] [CrossRef]
  25. Kaštil, J.; Kamarád, J.; Míšek, J.; Isnard, O.; Amara, M.; Arnold, Z. Magnetostriction and extraordinary exchange spring and exchange bias effects in Ni48Mn39Sn13 Heusler alloy. Intermetallics 2021, 132, 109137. [Google Scholar] [CrossRef]
  26. Huang, Y.; Qian, J.; Dong, D.; Shi, Y.; Du, Y.; Tang, S. Magnetostriction in <0kl>-oriented composites with CoMnSi microspheres. J. Magn. Magn. Mater. 2022, 543, 168621. [Google Scholar]
  27. Liua, J.; Gonga, Y.; Zhang, F.; Youa, Y.; Xu, G.; Miao, X.; Xu, F. Large, low-field and reversible magnetostrictive effect in MnCoSi-based metamagnet at room temperature. J. Mater. Sci. Technol. 2021, 76, 104–110. [Google Scholar] [CrossRef]
  28. Minorowicz, B.; Milecki, A. Design and Control of Magnetic Shape Memory Alloy Actuators. Materials 2022, 15, 4400. [Google Scholar] [CrossRef]
  29. Kurita, H.; Keino, T.; Senzaki, T.; Narita, F. Direct and inverse magnetostrictive properties of Fe–Co–V alloy particle-dispersed polyurethane matrix soft composite sheets. Sens. Actuators A Phys. 2022, 337, 113427. [Google Scholar] [CrossRef]
  30. Sadeghzadeh, A.; Asua, E.; Feuchtwanger, J.; Etxebarria, V.; García-Arribas, A. Ferromagnetic shape memory alloy actuator enabled for nanometric position control using hysteresis compensation. Sens. Actuators A Phys. 2012, 182, 122–129. [Google Scholar] [CrossRef]
  31. Kihara, T.; Roy, T.; Xu, X.; Miyake, A.; Tsujikawa, M.; Mitamura, H.; Tokunaga, M.; Adachi, Y.; Eto, T.; Kanomata, T. Observation of inverse magnetocaloric effect in magnetic-field-induced austenite phase of Heusler alloys Ni50−xCoxMn31.5Ga18.5 (x = 9 and 9.7). Phys. Rev. Mater. 2021, 5, 034416. [Google Scholar] [CrossRef]
  32. Hennel, M.; Galdun, L.; Džubinská, A.; Reiffers, M.; Varga, R. High efficiency direct magnetocaloric effect in Heusler Ni2MnGa microwire at low magnetic fields. J. Alloys Compd. 2023, 960, 170621. [Google Scholar] [CrossRef]
  33. Brock, J.; Khan, M. Large refrigeration capacities near room temperature in Ni2Mn1−xCrxIn. J. Magn. Magn. Mater. 2017, 425, 1–5. [Google Scholar] [CrossRef]
  34. Zheng, T.; Liu, K.; Chen, H.; Wang, C. Large magnetocaloric and magnetoresistance effects during martensitic transformation in Heusler-type Ni44Co6Mn37In13 alloy. J. Magn. Magn. Mater. 2022, 563, 170034. [Google Scholar] [CrossRef]
  35. Salazar-Mejía, C.; Devi, P.; Singh, S.; Felser, C.; Wosnitza, J. Influence of Cr substitution on the reversibility of the magnetocaloric effect in Ni-Cr-Mn-In Heusler alloys. Phys. Rev. Mater. 2021, 5, 104406. [Google Scholar] [CrossRef]
  36. Kitanovski, A. Energy applications of magnetocaloric materials. Adv. Energy Mater. 2020, 10, 1903741. [Google Scholar] [CrossRef]
  37. Sakon, T.; Nagashio, H.; Sasaki, K.; Susuga, S.; Numakura, D.; Abe, M.; Endo, K.; Nojiri, H.; Kanomata, T. Thermal expansion and magnetization studies of the novel ferromagnetic shape memory alloy N 2MnGa0.88Cu0.12 in a magnetic field. Phys. Scr. 2011, 84, 045603. [Google Scholar] [CrossRef]
  38. Khovailo, V.V.; Takagi, T.; Tani, J.; Levitin, R.Z.; Cherechukin, A.A.; Matsumoto, M.; Note, R. Magnetic properties of Ni2.18Mn0.82Ga heusler alloys with a coupled magnetostructural transition. Phys. Rev. B 2002, 65, 092410. [Google Scholar] [CrossRef] [Green Version]
  39. Filippov, D.A.; Khovailo, V.V.; Koledov, V.V.; Krasnoperov, E.P.; Levitin, R.Z.; Shavrov, V.G.; Takagi, T. The magnetic field influence on magnetostructural phase transition in Ni2.19Mn0.81Ga. J. Magn. Magn. Mater. 2003, 258, 507–509. [Google Scholar] [CrossRef] [Green Version]
  40. Khovailo, V.V.; Novosad, V.; Takagi, T.; Filippov, D.A.; Levitin, R.Z.; Vasil’ev, A.N. Magnetic properties and magnetostructural phase transitions in shape memory alloys. Phys. Rev. B 2004, 70, 174413. [Google Scholar] [CrossRef] [Green Version]
  41. Gonzàlez-Comas, A.; Obradó, E.; Mañosa, L.; Planes, A.; Chernenko, V.A.; Hattink, B.J.; Labarta, A. Premartensitic and martensitic phase transitions in ferromagnetic Ni2MnGa. Phys. Rev. B 1999, 60, 7085–7090. [Google Scholar] [CrossRef] [Green Version]
  42. Endo, K.; Kanomata, T.; Kimura, A.; Kataoka, M.; Nishihara, H.; Umetsu, R.Y.; Obara, K.; Shishido, T.; Nagasako, M.; Kainuma, R.; et al. Magnetic phase diagram of the ferromagnetic shape memory alloys Ni2MnGa1−xCux. Mater. Sci. Forum 2011, 684, 165–176. [Google Scholar] [CrossRef]
  43. Kanomata, T.; Endo, K.; Kudo, N.; Umetsu, R.Y.; Nishihara, H.; Kataoka, M.; Nagasako, M.; Kainuma, R.; Ziebeck, K.R.A. Magnetic moment of Cu-modified Ni2MnGa magnetic shape memory alloys. Metals 2013, 3, 114–122. [Google Scholar] [CrossRef] [Green Version]
  44. Kittel, C. Introduction of Solid State Physics, 8th ed.; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2004; p. 75. [Google Scholar]
  45. Singh, S.; Bednarcik, J.; Barman, S.R.; Felser, C.; Pandey, D. Premartensite to martensite transition and its implications for the origin of modulation in Ni2MnGa ferromagnetic shape memory alloy. Phys. Rev. B 2015, 92, 054112. [Google Scholar] [CrossRef] [Green Version]
  46. Nizhankovskii, V.I. Classical magnetostriction of nickel in high magnetic field. Eur. Phys. J. B 2006, 53, 1–4. [Google Scholar] [CrossRef]
  47. Wang, Z.; Liu, J.; Jiang, C.; Xu, H. The stress dependence of magnetostriction hysteresis in TbDyFe [110] oriented crystal. J. Appl. Phys. 2011, 109, 123923. [Google Scholar] [CrossRef]
  48. Sakon, T.; Matsumoto, T.; Komori, T. Rotation angle sensing system using magnetostrictive alloy Terfenol-D and permanent magnet. Sens. Actuators A 2021, 321, 112588. [Google Scholar] [CrossRef]
Figure 1. Temperature dependence of the magnetization, M vs. T, for Ni2MnGa0.88Cu0.12 in a magnetic field of 0.1 T. The arrows show the process of the temperature.
Figure 1. Temperature dependence of the magnetization, M vs. T, for Ni2MnGa0.88Cu0.12 in a magnetic field of 0.1 T. The arrows show the process of the temperature.
Metals 13 01185 g001
Figure 2. Differential magnetization at temperature, dM/dT vs. T, for Ni2MnGa0.88Cu0.12 in a magnetic field of 0.1 T. TM and TR indicate the martensitic transition temperature and the reverse martensitic transition temperature.
Figure 2. Differential magnetization at temperature, dM/dT vs. T, for Ni2MnGa0.88Cu0.12 in a magnetic field of 0.1 T. TM and TR indicate the martensitic transition temperature and the reverse martensitic transition temperature.
Metals 13 01185 g002
Figure 3. Magnetization process, M vs. H, for Ni2MnGa0.88Cu0.12 from 360 to 330 K with 2 K steps. The arrows show the process of the magnetic field.
Figure 3. Magnetization process, M vs. H, for Ni2MnGa0.88Cu0.12 from 360 to 330 K with 2 K steps. The arrows show the process of the magnetic field.
Metals 13 01185 g003
Figure 4. Magnetic field dependence of forced linear magnetostriction, ΔL/L vs. H, for Ni2MnGa0.88Cu0.12. The arrows show the process of the magnetic field.
Figure 4. Magnetic field dependence of forced linear magnetostriction, ΔL/L vs. H, for Ni2MnGa0.88Cu0.12. The arrows show the process of the magnetic field.
Metals 13 01185 g004
Figure 5. Magnetic field dependence of forced linear magnetostriction, ΔL/L vs. H, for Ni2MnGa0.88Cu0.12 up to 9.8 T. The arrows show the process of the magnetic field.
Figure 5. Magnetic field dependence of forced linear magnetostriction, ΔL/L vs. H, for Ni2MnGa0.88Cu0.12 up to 9.8 T. The arrows show the process of the magnetic field.
Metals 13 01185 g005
Figure 6. Magnetic field dependence of forced linear magnetostriction, ΔL/L vs. H, for Ni50Mn30Ga20. The arrows show the process of the magnetic field.
Figure 6. Magnetic field dependence of forced linear magnetostriction, ΔL/L vs. H, for Ni50Mn30Ga20. The arrows show the process of the magnetic field.
Metals 13 01185 g006
Table 1. Crystal structure, volume change, and linear magnetostriction ( Δ L / L )// value of the Ni2MnGa-type Heusler alloys. All alloys have the L21 cubic crystal structure in the parent austenite phase. Arrows indicate the transition direction of the magnetic and crystallographic phases.
Table 1. Crystal structure, volume change, and linear magnetostriction ( Δ L / L )// value of the Ni2MnGa-type Heusler alloys. All alloys have the L21 cubic crystal structure in the parent austenite phase. Arrows indicate the transition direction of the magnetic and crystallographic phases.
AlloyCrystal Structure
(Martensite Phase)
Volume Change (ppm)
(Paramagnetic Austenite
Ferromagnetic Martensite)
Linear Magnetostriction ( L / L (Paramagnetic Austenite
Ferromagnetic Martensite)
Ni2MnGa14M 1−330 1−780 2
Ni2MnGa0.88Cu0.12 3D022 + 14M−2000−1300 (this study)
Ni41Co9Mn31.5Ga18.5 4tetragonal57001900
Ni45Co5Mn36.7In13.3 514M---30,000 (single crystal)
Ni2Mn0.7Cu0.3Ga0.84
Al0.16 6
L10−55,000−26,000
Ni2.15Mn0.70Cr0.15Ga 7L10−30,000−8100
Ni50Mn30Ga20 85M−260−3300 (single crystal)
(this study)
1 Ref. [45] 2 Ref. [17] 3 Ref. [37] 4 Ref. [31] 5 Ref. [7] 6 Ref. [19] 7 Ref. [20] 8 Ref. [5]. The arrows indicate the transition direction of the magnetic and crystallographic phases.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sakon, T.; Morikawa, K.; Narumi, Y.; Hagiwara, M.; Kanomata, T.; Nojiri, H.; Adachi, Y. Magnetostriction of Heusler Ferromagnetic Alloy, Ni2MnGa0.88Cu0.12, around Martensitic Transition Temperature. Metals 2023, 13, 1185. https://doi.org/10.3390/met13071185

AMA Style

Sakon T, Morikawa K, Narumi Y, Hagiwara M, Kanomata T, Nojiri H, Adachi Y. Magnetostriction of Heusler Ferromagnetic Alloy, Ni2MnGa0.88Cu0.12, around Martensitic Transition Temperature. Metals. 2023; 13(7):1185. https://doi.org/10.3390/met13071185

Chicago/Turabian Style

Sakon, Takuo, Koki Morikawa, Yasuo Narumi, Masayuki Hagiwara, Takeshi Kanomata, Hiroyuki Nojiri, and Yoshiya Adachi. 2023. "Magnetostriction of Heusler Ferromagnetic Alloy, Ni2MnGa0.88Cu0.12, around Martensitic Transition Temperature" Metals 13, no. 7: 1185. https://doi.org/10.3390/met13071185

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop